You are on page 1of 15

chemosensors

Article
Influence of Nickel Doping on Ultrahigh Toluene Sensing
Performance of Core-Shell ZnO Microsphere Gas Sensor
Zhenhua Li 1 , Sijia Li 1 , Zijian Song 1 , Xueli Yang 1, * , Ziyan Wang 1 , Hao Zhang 2 , Lanlan Guo 3 , Caixuan Sun 1 ,
Hongyan Liu 1 , Junkai Shao 1 , Yehong Cheng 4, * and Guofeng Pan 1

1 Tianjin Key Laboratory of Electronic Materials and Devices, School of Electronics and Information
Engineering, Hebei University of Technology, 5340 Xiping Road, Beichen District, Tianjin 300401, China
2 Tianjin Key Laboratory of Materials Laminating Fabrication and Interface Control Technology,
School of Materials Science and Engineering, Hebei University of Technology, Tianjin 300130, China
3 School of Physics and Electronic Information Engineering, Henan Polytechnic University,
Jiaozuo 454003, China
4 School of Mechanical Engineering, Hebei University of Technology, Tianjin 300401, China
* Correspondence: xlyang@hebut.edu.cn (X.Y.); chengyehong@hebut.edu.cn (Y.C.)

Abstract: As a volatile organic compound, toluene is extremely harmful to the environment and
human health. In this work, through a simple one-step solvothermal method, Ni-doped ZnO
sensitive materials (0.5, 1, and 2 at% Ni-doped ZnO) with a core-shell morphology were synthesized
for the first time for toluene gas detection. The sensing test results showed that the sensor based
on 1 at% Ni-doped ZnO exhibited the best toluene sensing performance. The response was up
to 210 to 100 ppm toluene at 325 ◦ C. The sensor exhibited high selectivity, fast response/recovery
characteristics (2/77 s), and low detection limit (500 ppb, 3.5). Furthermore, we carried out molecular-
level research on the sensitive material prepared in this experiment by various characterization
Citation: Li, Z.; Li, S.; Song, Z.; Yang,
methods. The SEM characterization results showed that ZnO and Ni-doped ZnO possessed the core-
X.; Wang, Z.; Zhang, H.; Guo, L.; Sun,
C.; Liu, H.; Shao, J.; et al. Influence of shell morphology, and the average grain size decreased with the increase in the Ni doping content.
Nickel Doping on Ultrahigh Toluene The UV–Vis test showed that the band gap of ZnO became smaller with the increase in the Ni doping
Sensing Performance of Core-Shell amount. The enhanced toluene sensing performance of 1 at% Ni-doped ZnO could be ascribed to
ZnO Microsphere Gas Sensor. the structural sensitization and Ni doping sensitization, which are discussed in detail in the sensing
Chemosensors 2022, 10, 327. mechanism section.
https://doi.org/10.3390/
chemosensors10080327 Keywords: Ni doping; ZnO; gas sensor; toluene; core-shell structure
Academic Editors: Fanli Meng,
Zhenyu Yuan and Dan Meng

Received: 14 July 2022 1. Introduction


Accepted: 10 August 2022
With the development of modern industry, various kinds of poisonous and harmful
Published: 12 August 2022
gases have been produced, which will cause harm to human health and the environment.
Publisher’s Note: MDPI stays neutral In particular, volatile organic compounds (VOCs) such as ethanol, acetone, formaldehyde,
with regard to jurisdictional claims in and toluene, etc., which widely exist in our surroundings. Among them, toluene as a
published maps and institutional affil- typical aromatic compound is a kind of colorless gas with a special fragrance, and is
iations.
widely used in many aspects of production and life such as raw materials in the chemical
industry, the production of dyes, pharmaceuticals, pesticides, explosives, and other fine
chemicals. In addition, it can be used as a food preservative in major food production
factories [1]. Meanwhile, toluene is a flammable and harmful gas and can produce an
Copyright: © 2022 by the authors.
Licensee MDPI, Basel, Switzerland.
anesthetic effect on the central nervous system, causing great and irreparable harm to the
This article is an open access article
human body. Therefore, toluene has been identified as a highly toxic carcinogen by the
distributed under the terms and International Center for Research on Cancer [2]. Due to the harmfulness of toluene, it is
conditions of the Creative Commons very necessary and urgent to develop a high performance toluene gas sensor to monitor
Attribution (CC BY) license (https:// toluene in real-time to ensure human health and environmental safety.
creativecommons.org/licenses/by/ Gas sensors based on semiconductor metal oxides have attracted extensive attention in
4.0/). recent years due to their advantages of high sensitivity, easy fabrication, low cost, and small

Chemosensors 2022, 10, 327. https://doi.org/10.3390/chemosensors10080327 https://www.mdpi.com/journal/chemosensors


Chemosensors 2022, 10, 327 2 of 15

size. As the most promising sensing material, gas sensors based on semiconductor metal
oxides have become a research hotspot. Up to now, many metal oxide semiconductors have
been developed and used as gas sensing materials, for instance, ZnO [3,4], In2 O3 [5–7],
SnO2 [8–10], WO3 [11–13], NiO [14,15], CuO [16,17], etc. Among the n-type and p-type
single semiconductor oxides, ZnO is the most widely studied gas sensing material as
ZnO is an n-type semiconductor oxide with a wide band gap energy (3.37 eV) and a large
excite binding energy (60 MeV) [18,19]. The physical and chemical properties of ZnO are
stable, and the electrical property of ZnO is tunable with transition metal doping [20]. Cur-
rently, different morphologies of ZnO have been reported to detect different gases [21–27]
including nanoparticles [28], one-dimensional nanowires [29], nanorods and nanofibers,
two-dimensional nanosheets [30], nanoribbons and nanorings as well as three-dimensional
flower-like structures [31], core-shell structures, and hollow spheres. Studies have shown
that the hierarchical structure of ZnO enables better gas sensing properties due to its larger
specific surface area, porosity, and permeability.
In addition to constructing the hierarchical structure of ZnO, it is also an effective way
to improve the gas sensitive performance by doping. There are many types of doping sub-
stances including metal oxides, precious metals, transition metals, carbon-based materials,
and polymers [32–36]. Since transition metal ions can be doped into the semiconductor
lattice, the lattice distortion of sensitive materials generates a large number of defects and
vacancies. The increase in the defect concentration on the surface can improve the adsorp-
tion of oxygen anions on the surface of the material, thereby improving the gas-sensing
response value of the sensing material. Thus far, many studies have reported that through
transition metal doping, improved gas sensing performance can be achieved. Transition
metal doping elements including Co, Sn, Fe, and Ni [37–40] etc. have been reported to tune
the sensing performance of ZnO. The catalytic properties, surface properties, and charge
carriers of pristine semiconductor materials can be improved by doping transition metal
elements. Guo [41] et al. successfully synthesized Fe-doped ZnO/rGO nanocomposites
via a simple hydrothermal process. The sensor exhibited a response of 12.7 to 5 ppm
formaldehyde at 120 ◦ C. Compared with pure ZnO, the response of the doped sensing
material was about 3 times higher. The reason for the enhanced sensing performance is
that the doping of Fe changes the band gap, making the band gap of ZnO smaller and
generating a large number of oxygen vacancies.
Based on that demonstrated above, the gas sensing performance of semiconductor
metal oxides can be improved by constructing a hierarchical structure to adjust the surface
area and doping transition metal elements to tune the electrical and catalytic proper-
ties [42–44]. To our best knowledge, fabricating a core-shell Ni doped ZnO (Ni–ZnO) to
achieve an enhanced sensing performance of toluene has rarely been reported.
Here, materials with different ratios of 0, 0.5, 1, and 2 at% Ni–ZnO core-shell morpholo-
gies were successfully prepared by a simple solvothermal method and used for toluene
gas detection. Gas sensing measurements showed that out of all the four gas sensors, the
sensor based on 1 at% Ni–ZnO had the highest toluene response of 210 (100 ppm, 325 ◦ C),
which was almost seven times higher than pure ZnO (33.6, 375 ◦ C). Moreover, the sensor
also had high selectivity, low detection limit, and fast response/recovery properties. The
sensing mechanism of the enhanced sensing properties was analyzed in detail.

2. Materials and Methods


2.1. Synthesis of the Pure ZnO and Ni–ZnO Core-Shell Spheres
First, 20 mL of deionized water and 1 mmol Ni (CH3 COO)2 ·4H2 O were added into a
50 mL beaker and stirred for 20 min. Then, 7 mL of glycerol and 30 mL of isopropanol were
added into a 50 mL beaker with vigorous stirring, and 0.45 mmol Zn (CH3 COO)2 ·2H2 O
was dissolved into the above mixed solution with stirring for 15 min, named as solution
A. Then 0, 45, 90, and 180 µL of the above prepared Ni (CH3 COO)2 ·4H2 O solution were
added into solution A, respectively, and stirring was continued for another 20 min. Then,
the four solutions were placed in an oven and kept at 180 ◦ C for 24 h. After the reaction was
Chemosensors 2022, 10, 327 3 of 15

completed, the temperature was naturally cooled to room temperature, and the product
was separated by centrifugation, alternately washed with anhydrous ethanol and deionized
water, and then dried at 80 ◦ C for 10 h. The dried product was calcined at 500 ◦ C for 2 h
(heating rate, 2 ◦ C·min−1 ). Finally, pure ZnO and 0.5, 1, and 2 at% Ni–ZnO core-shell
products were obtained.

2.2. Characterization
In this paper, an X-ray powder diffractometer (XRD; Bruker D8 Discover, Billerica, MA,
USA) was used to investigate the composition and crystallinity of pure ZnO and different
proportions of the Ni–ZnO samples. Diffraction analysis of sensitive materials was carried
out using Cu Kα1 rays (λ = 1.5406 Å) in the range of 2θ from 20 to 80◦ , and the scanning step
was 10◦ ·min−1 . The microscopic topography of sensitive materials can be analyzed and
studied using scanning electron microscopy (SEM; ZEISS Sigma 500, Jena, Germany). More
information about the microscopic morphology and the lattice size of sensitive materials can
be obtained by transmission electron microscopy and high-resolution transmission electron
microscopy. In addition, energy dispersive X-ray spectrometry (EDS) enables the detection
of element mapping distributions. The chemical element analysis was obtained from the
X-ray photoelectron spectroscopy (XPS; Escalab 250 XI, Waltham, MA, USA) measurements,
and the surface area and pore size distribution were calculated by the Brunauer–Emmett–
Teller (BET) and Barrett–Joyner–Halenda (BJH) methods using the nitrogen adsorption–
desorption isotherm test (Micromeritics ASAP2000 system, USA). The absorbance and
band gap of the sample were measured by using a UV–Visible spectrophotometer (UV–Vis;
Lambda 1050+, Waltham, MA, USA) between 250 and 800 nm.

2.3. Fabrication and Measurement Process of the Gas Sensors


As schematically shown in Figure S1a,b, first, the as-synthesized sample (3–5 mg) and
an appropriate amount (0.3–0.5 mL) of deionized water were mixed to form a slurry. Then,
the slurry was evenly coated on the surface of the ceramic tube (outer diameter 1.2 mm,
inner diameter 0.8 mm, length 4 mm) using a small brush. After drying in air for 15 min,
the ceramic tube was placed in a Muffle furnace at 400 ◦ C for 120 min with a heating rate
of 2 ◦ C min−1 to enhance the stability during the test process. After that, a Ni–Cr alloy
wire, which provides the operating temperature by tuning the constant current, is inserted
in the sintered ceramic tube. Finally, the sensor was welded on a six-legged sensor base,
and the gas sensing performance was tested after aging for 24 h with 100 mA. Figure S1c
is schematically shown as the static gas sensing test system. A constant-current power
(Gwinstek GPD-3303S, New Taipei City, Taiwan) was used to supply a constant current
for the Ni–Cr alloy wire to control the operating temperature. Two glass cavities with a
volume of 1 L were used to hold air and test gas, respectively. A digital multimeter (Fluke
8846a, Everett, WA, USA) and a computer was used to record and monitor the resistance
of the gas sensor. During the sensing test process, the sensor was alternately placed in
1 L of fresh air in the glass cavity for a few minutes to obtain a stabilized resistance value
recorded as Ra. Then, the gas sensor device was transferred to another glass cavity with
1 L, which was filled with a mixture of fresh air and target gas, the stabilized resistance
value in the target gas was recorded as Rg. The sensing test was carried out in a laboratory
environment (25 ◦ C, 30 RH%). The volatile organic gas used was an analytical grade liquid,
and injected to the glass cavity by a microinjector. The gas preparation process is detailed
in the Supplementary Materials. The gas sensing response was defined as S = Ra/Rg,
and response/recovery time was the time taken by the sensor to reach 90% of the overall
resistance change.

3. Results
The XRD pattern of the pure ZnO, 0.5, 1, and 2 at% Ni–ZnO core-shell spheres are
shown in Figure 1a. The diffraction peaks appeared at 34.42◦ , 36.25◦ , 47.53◦ , 56.60◦ , 62.85◦ ,
68.17◦ , 69.29◦ , which corresponded to the (100), (002), (101), (102), (110), (103), and (112)
3. Results

Chemosensors 2022, 10, 327


The XRD pattern of the pure ZnO, 0.5, 1, and 2 at% Ni–ZnO core-shell spheres4 are of 15
shown in Figure 1a. The diffraction peaks appeared at 34.42°, 36.25°, 47.53°, 56.60°, 62.85°,
68.17°, 69.29°, which corresponded to the (100), (002), (101), (102), (110), (103), and (112)
crystal planes of the hexagonal wurtzite phase ZnO (JCPDS No. 36-1451) [45]. However,
crystal planes of the hexagonal wurtzite phase ZnO (JCPDS No. 36-1451) [45]. However,
no peak matched with Ni or NiO in the Ni–ZnO samples. No other diffraction peaks cor-
no peak matched with Ni or NiO in the Ni–ZnO samples. No other diffraction peaks
responding to the impurities were detected. Figure 1b is the high-resolution peak of the
corresponding to the impurities were detected. Figure 1b is the high-resolution peak of the
(101) plane of ZnO compared to pure ZnO, and the diffraction peak of Ni–ZnO shifted to
(101) plane of ZnO compared to pure ZnO, and the diffraction peak of Ni–ZnO shifted to a
a higher angle direction. The shift of 2 at% Ni–ZnO sample was the most obvious, which
higher angle direction. The shift of 2 at% Ni–ZnO sample was the most obvious, which
indicated that Ni2+2+was incorporated in the ZnO crystal lattice due to the radius of Ni2+2+
indicated that Ni was incorporated in the ZnO crystal lattice due to the radius of Ni
being smaller than Zn2+2+
.
being smaller than Zn .

(a)The
Figure1.1.(a)
Figure TheXRD
XRDpatterns
patternsofofthe
thepure
pureand
andNi–ZnO,
Ni–ZnO,(b)
(b)enlarged
enlargedspectra
spectraofofthe
theXRD
XRDpatterns
patternsofof
the
theNi–ZnO.
Ni–ZnO.

Figure2a–d
Figure 2a–dshows
showsthe the
SEMSEM images
images of the
of the 0, 0.5,
0, 0.5, 1, and1, and
2 at%2 Ni–ZnO
at% Ni–ZnO sample,
sample, re-
respec-
spectively. The pure ZnO, 0.5, and 1 at% Ni–ZnO samples showed
tively. The pure ZnO, 0.5, and 1 at% Ni–ZnO samples showed a core-shell spherical hier- a core-shell spherical
hierarchical
archical structure,
structure, and there
and there was no was no obvious
obvious core-shell
core-shell structure
structure observed
observed for 2 for
at%2Ni–at%
Ni–ZnO;
ZnO; the morphology
the morphology was awassimplea simple spherical
spherical structure.
structure. It wasfrom
It was found found thefrom
SEMthe SEM
images
images
that withthat
thewith the increase
increase in the Niin doping
the Ni doping
content,content, the surface
the surface of theofZnO the ZnO microspheres
microspheres be-
became more and more coarse and the average particle size of the
came more and more coarse and the average particle size of the spheres decreased. The spheres decreased. The
average particle size was 15.10, 9.42, 5.62, and 4.66 µm for pure ZnO,
average particle size was 15.10, 9.42, 5.62, and 4.66 μm for pure ZnO, 0.5, 1 and 2 at% Ni– 0.5, 1 and 2 at%
Ni–ZnO spheres, which suggests that the addition of Ni ions can
ZnO spheres, which suggests that the addition of Ni ions can inhibit the growth of ZnOinhibit the growth of ZnO
particlesin
particles inthe
thehydrothermal
hydrothermalreactionreactionprocess.
process.In Inaddition,
addition,the theinserted
insertedmagnified
magnifiedSEM SEM
imagesshowed
images showedthat thatthe
thecore
coreofofthe
thecore-shell
core-shellspheres
sphereswaswasvery verylarge
largeandandnext
nexttotothe
theshell.
shell.
Proper Ni doping, a rough surface, and loose shell are favorable
Proper Ni doping, a rough surface, and loose shell are favorable for gas adsorption and for gas adsorption and
diffusion, with a high gas sensing performance, which will be verified
diffusion, with a high gas sensing performance, which will be verified in later sections. in later sections.
Figure 3a,b are TEM images of different magnifications, whose shape and size are
consistent with the SEM characterization results. The inset figure is the selected area electron
diffraction (SAED) image and shows that the 1 at% Ni–ZnO sample had a polycrystalline
structure. Figure 3c is the HRTEM images of the 1 at% Ni–ZnO sample. The lattice spacings
were 0.286 and 0.249 nm, corresponding to the (100) and (101) crystal planes of ZnO [46,47].
Interplanar spacing corresponding to NiO was not found, which is consistent with the
results of the XRD. From the element mapping test results in Figure 3d–g, the three elements
(Ni, O, and Zn) were evenly distributed in the 1 at% Ni–ZnO sample.
The specific surface area of the pure ZnO and Ni–ZnO samples was determined by
the nitrogen adsorption/desorption isotherm. As illustrated in Figure 4a–e, the adsorption–
desorption curve is an iv-type isotherm belonging to the H3 hysteresis loop, indicating
that Ni–ZnO has a microporous structure [48].The specific surface area increased with
the increase in the Ni doping amount, especially for 1 and 2 at% Ni–ZnO, the specific
surface areas were 20.43 to 26.41, 37.54 and 40.14 m2 ·g−1 for the pure ZnO, 0.5, 1, and 2 at%
Ni–ZnO. The results showed that with the proper amount of Ni doping, the specific area of
ZnO could increase a lot; when the doping amount reached a certain value, the specific area
did not increase too much. A large specific area could supply more surface-active sites for
2022, 10, Chemosensors
x FOR PEER2022, 10, 327
REVIEW 5 of 16 5 of 15

surface adsorbed oxygen species, thus improving the recognition function of the sensing
material and contributing to the high gas sensing performance.
Moreover, the inset of Figure 4 shows the nonlocal density functional theory (NLDFT)
and grand canonical Monte Carlo (GCMC) pore size distribution of the Ni–ZnO, and
through analysis, it was found that the pore sizes of the pure ZnO and 0.5 at%, 1 at%, and 2
at% Ni–ZnO were 46.8, 58.08, 32.72, and 32.71 nm, respectively. Appropriate Ni ion doping
can increase the size of the pore size, and combined with the gas sensing data, it can be
s 2022, 10, x FOR PEER REVIEW inferred that the pore size of the sensitive material in this experiment had little effect
5 of 16 on the
gas sensing performance. The increase in the specific surface area is one of the main factors
affecting the gas sensing performance.

Figure 2. The SEM images of the (a) pure (b) 0.5 at% (c) 1 at%, and (d) 2 at% Ni–ZnO.

Figure 3a,b are TEM images of different magnifications, whose shape and size are
consistent with the SEM characterization results. The inset figure is the selected area elec-
tron diffraction (SAED) image and shows that the 1 at% Ni–ZnO sample had a polycrys-
talline structure. Figure 3c is the HRTEM images of the 1 at% Ni–ZnO sample. The lattice
spacings were 0.286 and 0.249 nm, corresponding to the (100) and (101) crystal planes of
ZnO [46,47]. Interplanar spacing corresponding to NiO was not found, which is consistent
with the results of the XRD. From the element mapping test results in Figure 3d–g, the
three
Figureelements (Ni,
2. The SEM O, and
images of Zn) were
the (a) pureevenly
(b) 0.5 distributed
at% (c) 1 at%,inand
the(d)
1 at%
2 at%Ni–ZnO
Ni–ZnO.sample.
Figure 2. The SEM images of the (a) pure (b) 0.5 at% (c) 1 at%, and (d) 2 at% Ni–ZnO.

Figure 3a,b are TEM images of different magnifications, whose shape and size are
consistent with the SEM characterization results. The inset figure is the selected area elec-
tron diffraction (SAED) image and shows that the 1 at% Ni–ZnO sample had a polycrys-
talline structure. Figure 3c is the HRTEM images of the 1 at% Ni–ZnO sample. The lattice
spacings were 0.286 and 0.249 nm, corresponding to the (100) and (101) crystal planes of
ZnO [46,47]. Interplanar spacing corresponding to NiO was not found, which is consistent
with the results of the XRD. From the element mapping test results in Figure 3d–g, the
three elements (Ni, O, and Zn) were evenly distributed in the 1 at% Ni–ZnO sample.

Figure 3. The TEM image


Figure of 1TEM
3. The at%image
Ni–ZnOof 1 (a,b). (c) HRTEM
at% Ni–ZnO (a,b). images of 1images
(c) HRTEM at% Ni–ZnO. (d) The (d) The
of 1 at% Ni–ZnO.
overlapped elemental mapping
overlapped of Ni, O,
elemental Zn in the
mapping 1 at%
of Ni, O, Ni–ZnO
Zn in thesample, and (e–g)
1 at% Ni–ZnO elemental
sample, map-elemental
and (e–g)
ping of Ni, O, andmapping
Zn. of Ni, O, and Zn.

The specific surface area of the pure ZnO and Ni–ZnO samples was determined by
the nitrogen adsorption/desorption isotherm. As illustrated in Figure 4a–e, the adsorp-
tion–desorption curve is an iv-type isotherm belonging to the H3 hysteresis loop, indicat-
ing that Ni–ZnO has a microporous structure [48].The specific surface area increased with
the increase in the Ni doping amount, especially for 1 and 2 at% Ni–ZnO, the specific
(NLDFT) and grand canonical Monte Carlo (GCMC) pore size distribution of the Ni–ZnO,
and through analysis, it was found that the pore sizes of the pure ZnO and 0.5 at%, 1 at%,
and 2 at% Ni–ZnO were 46.8, 58.08, 32.72, and 32.71 nm, respectively. Appropriate Ni ion
doping can increase the size of the pore size, and combined with the gas sensing data, it
can be inferred that the pore size of the sensitive material in this experiment had little
Chemosensors 2022, 10, 327 6 of 15
effect on the gas sensing performance. The increase in the specific surface area is one of
the main factors affecting the gas sensing performance.

Figure 4. Thenitrogen
4. The nitrogen absorption–desorption
absorption–desorption isotherms
isotherms and pore
and pore size distribution
size distribution curves curves (inset)
(inset) of
the (a) pure ZnO and (b) 0.5 at% (c) 1 at%, and (d) 2 at% Ni–ZnO. (e) Specific surface area
of the (a) pure ZnO and (b) 0.5 at% (c) 1 at%, and (d) 2 at% Ni–ZnO. (e) Specific surface area of of the
samples.
the samples.

The XPS measurement was used to analyze the chemical composition and states of
the elements in the pure ZnO and Ni–ZnO samples. Figure 5a exhibits the full spectrum
of the ZnO and Ni–ZnO samples, where the full spectrum contained the peaks of Zn, O
and Ni, proving the presence of Zn, O, and Ni in the sample. In Figure 5b, the binding
energy around 1021.20 and 1044.28 eV corresponded to Zn 2p3/2 and Zn 2p1/2 , respectively,
proving that Zn was in the form of Zn2+ [49]. Figure 5c–e shows that the Ni 2p3/2 orbital
can be separated into three peaks of about 854.1, 856.3, and 861.2 eV, corresponding to the
Ni2+ , Ni3+ and satellite peaks, respectively [50], indicating the existence of Ni ions in the
ZnO crystals.
Figure 6a–d is the XPS spectra of the O 1s of all samples, the peak shape is asymmetric,
and can be allocated into three different oxygen components: chemisorbed gen (OC ),
oxygen vacancy (OV ), and lattice oxygen (OL ), which correspond to the peaks at around
532.2 ± 0.5 eV, 531.0 ± 0.5 eV, and 529.5 ± 0.5 eV [51,52], respectively. Table 1 shows the
relative proportion of the three oxygen components of O 1s in all of the samples. The
relative ratio of OC did not change a lot, but with Ni doping, the ratio of OL decreased,
while relatively, the proportion of OV increased, and 1 at% Ni–ZnO showed the highest
The XPS measurement was used to analyze the chemical composition and states of
the elements in the pure ZnO and Ni–ZnO samples. Figure 5a exhibits the full spectrum
of the ZnO and Ni–ZnO samples, where the full spectrum contained the peaks of Zn, O
Chemosensors 2022, 10, 327 and Ni, proving the presence of Zn, O, and Ni in the sample. In Figure 5b, the binding 7 of 15
energy around 1021.20 and 1044.28 eV corresponded to Zn 2p3/2 and Zn 2p1/2, respectively,
proving that Zn was in the form of Zn2+ [49]. Figure 5c–e shows that the Ni 2p3/2 orbital
can be separated
proportion of OV into
withthree peaks
22.8%. of about
Ni doping led854.1,
to the856.3, and 861.2
formation of aneV, corresponding
oxygen vacancy intoZnO,
the
Ni 2+, Ni3+ and satellite peaks, respectively [50], indicating the existence of Ni ions in the
and more OV concentration distributed to more surface-active sites, thus benefiting more
ZnO
surfacecrystals.
negatively charged oxygen species.

Chemosensors 2022, 10, x FOR PEER REVIEW 8 of 16


Figure 5. The
Figure The XPS spectra of the pure and Ni–ZnO for (a) the
the full
full scan,
scan, (b)
(b) Zn
Zn 2p,
2p,(c–e)
(c–e)Ni
Ni2p
2p3/2 of
3/2 of

0.5,
0.5, 1,
1, 22 at%
at% Ni–ZnO.
Ni–ZnO.

Figure 6a–d is the XPS spectra of the O 1s of all samples, the peak shape is asymmet-
ric, and can be allocated into three different oxygen components: chemisorbed gen (OC),
oxygen vacancy (OV), and lattice oxygen (OL), which correspond to the peaks at around
532.2 ± 0.5 eV, 531.0 ± 0.5 eV, and 529.5 ± 0.5 eV [51,52], respectively. Table 1 shows the
relative proportion of the three oxygen components of O 1s in all of the samples. The rel-
ative ratio of OC did not change a lot, but with Ni doping, the ratio of OL decreased, while
relatively, the proportion of OV increased, and 1 at% Ni–ZnO showed the highest propor-
tion of OV with 22.8%. Ni doping led to the formation of an oxygen vacancy in ZnO, and
more OV concentration distributed to more surface-active sites, thus benefiting more sur-
face negatively charged oxygen species.

Table 1. The relative percentages of O 1s in the XPS spectra ZnO, 0.5 at%, 1 at%, and 2 at% Ni–
ZnO.

Species Peak (eV) ZnO 0.5 at% Ni–ZnO 1 at% Ni–ZnO 2 at% Ni–ZnO
OL 529.9 60.0% 55.9% 53.0% 53.7%
OV 530.8 15.6% 19.9% 22.8% 21.0%
OC 531.7 24.4% 24.2% 24.2% 25.3%

Figure
Figure6.6.The
TheXPS
XPSspectra
spectraof
ofO
O1s
1sfor
for(a–d)
(a–d)pure,
pure,0.5,
0.5, 1,
1, and
and 22 at%
at% Ni–ZnO.
Ni–ZnO.

Gas Sensing Characteristics


The operating temperature can affect the carrier concentration and gas adsorption
and desorption of the sensor resistance and gas response [53,54]. Response of the four
sensors to 100 ppm toluene at different temperatures (Figure 7a) showed that the response
values first increased and then decreased. The maximum response values of the four sen-
Chemosensors 2022, 10, 327 8 of 15

Table 1. The relative percentages of O 1s in the XPS spectra ZnO, 0.5 at%, 1 at%, and 2 at% Ni–ZnO.

Species Peak (eV) ZnO 0.5 at% Ni–ZnO 1 at% Ni–ZnO 2 at% Ni–ZnO
OL 529.9 60.0% 55.9% 53.0% 53.7%
OV 530.8 15.6% 19.9% 22.8% 21.0%
OC 531.7 24.4% 24.2% 24.2% 25.3%

Gas Sensing Characteristics


The operating temperature can affect the carrier concentration and gas adsorption and
desorption of the sensor resistance and gas response [53,54]. Response of the four sensors to
100 ppm toluene at different temperatures (Figure 7a) showed that the response values first
increased and then decreased. The maximum response values of the four sensors reached
at 375 ◦ C (pure ZnO), 350 ◦ C (0.5 at% Ni–ZnO), and 325 ◦ C (1 and 2 at% Ni–ZnO), and the
response values were 33.6, 111.4, 210.0 and 23.1, respectively. The 1 at% Ni–ZnO sample
had the highest response of 210.0, which was seven times higher than the pure ZnO. With
Ni doping, the operating temperature of 1 at% Ni–ZnO was significantly decreased, which
could be explained by the change in the band gap, as illustrated in Figure S2. After Ni
doping, the band gap reduced from 3.05 to 2.89 eV, and free electrons were easier to release
from the conduction band. Figure 7b displays the responses of the four sensors to the six
tested target gases (toluene, ethanol, acetone, methanol, formaldehyde, and xylene) with
100 ppm at their optimal operating temperatures (ZnO, 375 ◦ C; 0.5 at% Ni–ZnO, 350 ◦ C;
1 and 2 at% Ni–ZnO, 325 ◦ C). Obviously, the four sensors had the highest response to
toluene. Compared with the other tested gases, the 1 at% Ni–ZnO sensor showed the best
toluene sensing performance, and the response to toluene and other gases was 210.0, 52.0, 9 of 16
Chemosensors 2022, 10, x FOR PEER REVIEW
3.4, 9.4, 3.1, and 1.5. The toluene response was 4–140 times higher than other measured
gases, indicating that the sensor had excellent selectivity for toluene.

Figure
Figure7.7. (a)
(a)The
Theresponse
response of the
the four
four gas
gassensors
sensorstoto100
100ppm
ppmtoluene
toluene
at at different
different temperatures. (b)
temperatures.
The selectivity of the four gas sensors to 100 ppm different VOC gases at their optimum
(b) The selectivity of the four gas sensors to 100 ppm different VOC gases at their operating
optimum
temperature.
operating temperature.

Figure 8a,b shows the real-time response curves of the four sensors at 375 C (pure ◦
Figure 8a,b shows the real-time response curves of the four sensors at 375 °C (pure
ZnO) and 325 ◦ C (Ni–ZnO samples) with different concentrations of toluene. The response
ZnO) and 325 °C (Ni–ZnO samples) with different concentrations of toluene. The response
value increased with the increase in toluene concentration. The sensor response based
value
on 1 increased
at% Ni–ZnO with the increase
increased in toluene
significantly withconcentration. The sensor
increasing toluene responseThe
concentration. based on
1corresponding
at% Ni–ZnOlinearincreased
relationship between the toluene response and concentration is listed corre-
significantly with increasing toluene concentration. The
sponding linear
in Figure 8c,d. relationship
It can be observed between
that the the toluene
linearity response
of the andvery
sensor was concentration
good in bothis listed in
the
Figure 8c,d. It canrange
low-concentration be observed that the linearityrange.
and the high-concentration of theThe
sensor was increased
response very good in both the
almost
linearly with the toluene
low-concentration range andconcentration, and 1 at% Ni–ZnO
the high-concentration range.had
Thethe highest increased
response sensitivity.almost
linearly with the toluene concentration, and 1 at% Ni–ZnO had the highest sensitivity.
Furthermore, the sensor based on 1 at% Ni–ZnO had a low detection limit of 0.5 ppm with
a response value of 3.5, which makes it promising in practical application.
ZnO) and 325 °C (Ni–ZnO samples) with different concentrations of toluene. The response
value increased with the increase in toluene concentration. The sensor response based on
1 at% Ni–ZnO increased significantly with increasing toluene concentration. The corre-
sponding linear relationship between the toluene response and concentration is listed in
Chemosensors 2022, 10, 327 Figure 8c,d. It can be observed that the linearity of the sensor was very good in 9both of 15 the
low-concentration range and the high-concentration range. The response increased almost
linearly with the toluene concentration, and 1 at% Ni–ZnO had the highest sensitivity.
Furthermore, the
Furthermore, thesensor
sensorbased
basedon
on1 1at%
at%Ni–ZnO
Ni–ZnO had
hada low detection
a low limit
detection of 0.5
limit ppm
of 0.5 withwith
ppm
aa response
response value
value of
of3.5,
3.5,which
whichmakes
makesit itpromising
promisingin in
practical application.
practical application.

Figure 8.
Figure (a,b) The
8. (a,b) The dynamic
dynamicresponse
responsecurves
curvesofofthe
thefour sensors
four to to
sensors toluene with
toluene different
with concentra-
different concentra-
tions at 375 ◦ C (pure ZnO) and 325 ◦ C (Ni–ZnO samples). (c,d) The response of the pure, 0.5, 1, and
tions at 375 °C (pure ZnO) and 325 °C (Ni–ZnO samples). (c,d) The response of the pure, 0.5, 1, and
22at%
at% Ni–ZnO
Ni–ZnO sensors
sensorsto
to(0.5–100
(0.5–100ppm)
ppm)toluene.
toluene.

The response and recovery characteristics are also important parameters of gas sensors.
Thus, the response and recovery transients of the sensor based on 1 at% Ni–ZnO to 100 ppm
toluene at 325 ◦ C are shown in Figure 9a. The response time and recovery time of the
sensor based on 1 at% Ni–ZnO are 2 and 77 s, respectively, illustrating that the sensor had a
relatively fast response and recovery property. Table 2 is a comparison of the toluene sensing
performance of this work and the sensors reported in the literature [55–59]. The operating
temperature was relatively higher than that in the literature, but the response and recovery
time were fast and the toluene response was much higher than the reported literature.
Figure 9b is the resistance in air as a function of operating temperature. With the
increase in Ni doping, the resistance of the sensor increased significantly. The resistance
versus temperature plot of the sensors showed an abnormal PTCR (positive temperature
coefficient of resistance) behavior during 300–350 ◦ C. Based on the reported literature [60],
this phenomenon was caused by Ni doping, which introduced defects in ZnO, thus forming
oxygen vacancy-like defects in this temperature region.
Figure 9c is a seven-cycle test curve of the 1 at% Ni–ZnO sample to 100 ppm toluene
at 325 ◦ C. The resistance in air and toluene was within the allowable fluctuation range,
proving that the sensor had good repeatability. The long-term stability was also studied in
this work; the sensing response to 100 ppm toluene in the air for one month is shown in
Figure 9d. During the test days, the response values varied and slightly decreased but did
not show an obvious fluctuation. The results indicate that the 1 at% Ni–ZnO gas sensor has
good long-term stability.
Figure 9c is a seven-cycle test curve of the 1 at% Ni–ZnO sample to 100 ppm toluene
at 325 °C. The resistance in air and toluene was within the allowable fluctuation range,
proving that the sensor had good repeatability. The long-term stability was also studied
in this work; the sensing response to 100 ppm toluene in the air for one month is shown
in Figure 9d. During the test days, the response values varied and slightly decreased but
Chemosensors 2022, 10, 327 did not show an obvious fluctuation. The results indicate that the 1 at% Ni–ZnO gas10 sen-
of 15

sor has good long-term stability.

Figure 9. (a) The


9. (a) The response/recovery
response/recoverytime timeofof11at%
at%Ni–ZnO
Ni–ZnOgas gassensor
sensor to
to 100
100 ppm
ppm toluene
toluene at
at 325 ◦ C.
325 °C.
Figure
(b) Resistance
(b) Resistanceininthe
theairair
ofof
thethe
four gasgas
four sensors at different
sensors operating
at different temperatures.
operating (c) Seven
temperatures. reversible
(c) Seven re-
cycles ofcycles
versible 1 at% of
Ni–ZnO gas sensors.
1 at% Ni–ZnO gas (d) The stability
sensors. (d) The testing
stabilitycurves ofcurves
testing 1 at% Ni–ZnO
of 1 at% gas sensorgas
Ni–ZnO to
sensor
100 ppmto toluene
100 ppmintoluene
30 days.in 30 days.

Table 2. A comparison of the toluene gas sensing performance of this work and other sensing material
in previously reported works.

Operating
Sensing Material Conc. in ppm Response Res/Rec (s) LOD (ppm) Ref.
Temperature (◦ C)
WO3 –SnO2 10 340 5.6 14.5/406 - [55]
SnO2 –ZnO 50 200 7.5 90/150 0.1 [56]
NiGa2 O4 –NiO 100 230 12.7 60/70 0.5 [57]
NiO/NiGa2 O4 5 200 10.54 600/- - [58]
CuO–SnO2 75 400 540 100/36 10 [59]
Ni–ZnO 100 320 210 2/77 0.5 Present work

4. Discussion
The gas sensing mechanism can be explained by the interaction of multiple factors.
In this work, toluene sensing mechanisms with exceptional response are discussed in
detail. The gas sensing mechanism has been widely explored in many studies [61,62],
and the extensively accepted theory is the chemical reaction between the target gas and
chemically adsorbed oxygen species (O2 − (T < 100 ◦ C), O− (100 ◦ C < T < 300 ◦ C), and O2−
(T > 300 ◦ C)) [63,64]. When redox reactions take place on the surface of the semiconduc-
tors, charge transfer will occur during this process, which changes the resistance of the
gas sensor.
As illustrated in Figure 10, when the Ni–ZnO gas sensor is exposed to air, O2 molecules
will capture electrons in the conduction band of the material to form negatively charged
oxygen species on the surface of the Ni–ZnO material, and a depletion layer will form on
the surface at the same time, leading to a high resistance of the sensor in air. Then, when
Chemosensors 2022, 10, x FOR PEER REVIEW 12 of 16

Chemosensors 2022, 10, 327 11 of 15


Ni–ZnO and the sensing performance will be enhanced. The third and most important
point is that Ni2+ ions can be easily oxidized into Ni3+ with a higher oxidation state, which
will
thefacilitate
sensor isthe redox to
exposed reaction, due gas
a reducing to which
(toluene)Ni3+environment,
usually playsthe thechemically
role of the adsorbed
catalyst
during the redox process, as many studies have reported [70–72]. Therefore,
oxygen can react with the target gas molecules on the surface of the sensing material, a greater per-
the
centage of Ni 3+ will obtain a good gas sensing property to a large extent. Figure 5c–e dis-
electron depletion layer becomes smaller. As a result, electrons that are released back
plays the conduction
into the fitted XPS results of Ni2+ lower
band induce and Niresistance
3+ of the 0.5, 1, and 2 at% Ni–ZnO. The relative
values in reducing gases. The relevant
percentage of Ni
reactions are seen in Equations (1)–(3) (T > 300 ◦ C) [65]:and the ratio of Ni /Ni was cal-
2+ and Ni3+ varied in the three samples, 3+ 2+

culated to be 1.33, 1.62, and 1.05 for the 0.5, 1, and 2 at% Ni–ZnO samples. The 1 at% Ni–
ZnO sample had the highest ratio of Ni3+O /Ni
2 → of
2+ 1.62, followed by the 0.5 at% Ni–ZnO and
O2(ads) (1)
2 at% Ni–ZnO. From the gas sensing tests, the highest sensing performance was obtained
2−
by the 1 at% Ni–ZnO, and then the 0.5 O2(ads) 4e → 2Oand
at%+Ni–ZnO 2 at% Ni–ZnO samples. The gas (2)
sensing results coincided with the inference of the Ni 3+/Ni2+ ratio.
C7 H8 + 18O2− → 4H2 O + 7CO2 + 36e− (3)

Figure 10. A schematic diagram of the sensing mechanism.


Figure 10. A schematic diagram of the sensing mechanism.
The enhanced sensing performance of the 1 at% Ni–ZnO-based gas sensor can be
attributed to the following two aspects including structural sensitization and modified
Chemosensors 2022, 10, 327 12 of 15

sensitization. The first is structural sensitization. After Ni doping, the hierarchical mi-
crostructure of ZnO changed, and the surface of the Ni–ZnO core-shell spheres became
rougher. This loose structure is very favorable for the diffusion of gas molecules, which can
not only penetrate into the shell at the outer surface of the material, but may also further
diffuse to the surface of the core. Therefore, the sensing material has a higher sensing
performance [66]. Furthermore, the thickness of the shell of the sensing material becomes
thinner with increasing Ni content. According to the literature [67,68], when the thickness
of the shell layer is close to the Debye length, the gas sensing material has the best gas
sensing performance. Finally, the size of the topography of the gas sensing material is also
affected by Ni, resulting in a smaller morphology size. An appropriate reduction in the
morphology size can lead to an increase in the specific surface area, which can adsorb more
oxygen and toluene molecules and provide more active sites [69].
Modified sensitization includes the following points. First, the specific surface area
increased with Ni doping, so the recognition function of the sensing material will be
improved. Second, it can be seen from Table 2 that Ni doping increased the relative
proportion of oxygen vacancy in ZnO. The oxygen vacancy improved from 15.6% to 22.8
for 1 at% Ni–ZnO, which means that more oxygen species will adsorb on the surface of
1 at% Ni–ZnO and the sensing performance will be enhanced. The third and most important
point is that Ni2+ ions can be easily oxidized into Ni3+ with a higher oxidation state, which
will facilitate the redox reaction, due to which Ni3+ usually plays the role of the catalyst
during the redox process, as many studies have reported [70–72]. Therefore, a greater
percentage of Ni3+ will obtain a good gas sensing property to a large extent. Figure 5c–e
displays the fitted XPS results of Ni2+ and Ni3+ of the 0.5, 1, and 2 at% Ni–ZnO. The
relative percentage of Ni2+ and Ni3+ varied in the three samples, and the ratio of Ni3+ /Ni2+
was calculated to be 1.33, 1.62, and 1.05 for the 0.5, 1, and 2 at% Ni–ZnO samples. The
1 at% Ni–ZnO sample had the highest ratio of Ni3+ /Ni2+ of 1.62, followed by the 0.5 at%
Ni–ZnO and 2 at% Ni–ZnO. From the gas sensing tests, the highest sensing performance
was obtained by the 1 at% Ni–ZnO, and then the 0.5 at% Ni–ZnO and 2 at% Ni–ZnO
samples. The gas sensing results coincided with the inference of the Ni3+ /Ni2+ ratio.

5. Conclusions
In summary, different Ni–ZnO core-shell spheres were successfully prepared by a one-
step solvothermal method and gas sensors based on the prepared materials were prepared.
The role of Ni doping in ZnO on the microstructure and gas sensing performance was stud-
ied in detail. The results showed that Ni doping changed the hierarchical microstructure
of ZnO, and the BET specific area increased after Ni doping. Gas sensing measurements
revealed that the operating temperature decreased due to the lower band gap after Ni
doping. All of the sensing materials had the highest response to toluene, the best sensing
performance was acquired by the 1 at% Ni–ZnO based gas sensor, with high response,
excellent selectivity, fast response and recovery time, and relatively long-term stability.
The enhanced sensing property can be mainly due to the increase in the specific surface
area and relative percentage of oxygen vacancy after Ni doping, and more importantly,
is the catalyst effect of Ni ions. This work provides a reasonable way to fabricate a high
performance toluene sensing material.

Supplementary Materials: The following supporting information can be downloaded at: https://
www.mdpi.com/article/10.3390/chemosensors10080327/s1, Figure S1: The schematic figure of the
(a) ceramic tube, (b) sensor device, and (c) the sensing test system. Figure S2: (a) The UV–Vis
absorption spectrum and (b) energy band gap of the ZnO samples with different Ni doping amounts.
Table S1: The parameter information for all of the gas samples.
Chemosensors 2022, 10, 327 13 of 15

Author Contributions: Conceptualization, X.Y. and Z.L.; Methodology, X.Y.; Formal analysis, Z.L.;
Investigation, S.L., Z.S., Z.W., H.Z., C.S., H.L. and J.S.; Data curation, Z.L.; Writing—original draft
preparation, Z.L.; Writing—review and editing, X.Y., G.P., Y.C. and L.G.; Supervision, X.Y., G.P., Y.C.
and L.G.; Funding acquisition, X.Y. and G.P. All authors have read and agreed to the published
version of the manuscript.
Funding: This work was supported by the Major National Science and Technology Special Projects
(2016ZX02301003-004-007); the National Natural Science Foundation of China (62003123); the National
Natural Science Foundation of Hebei Province (F2020202067); the Key Laboratory of Electronic
Materials and Devices of Tianjin, China; and the National Demonstration Center for Experimental
(Electronic and Communication Engineering) Education (Hebei University of Technology).
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Not applicable.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Mirzaei, A.; Leonardi, S.G.; Neri, G. Detection of hazardous volatile organic compounds (VOCs) by metal oxide nanostructures-
based gas sensors: A review. Ceram. Int. 2016, 42, 15119–15141. [CrossRef]
2. Xu, J.M.; Cheng, J.P. The advances of Co3 O4 as gas sensing materials: A review. J. Alloys Compd. 2016, 686, 753–768. [CrossRef]
3. Sharma, B.; Sharma, A.; Joshi, M.; Myung, J.H. Sputtered SnO2 /ZnO heterostructures for improved NO2 gas sensing properties.
Chemosensors 2020, 8, 67. [CrossRef]
4. Hui, G.Z.; Zhu, M.Y.; Yang, X.L.; Liu, J.J.; Pan, G.F.; Wang, Z.Y. Highly sensitive ethanol gas sensor based on CeO2 /ZnO binary
heterojunction composite. Mater. Lett. 2020, 278, 128453. [CrossRef]
5. Guo, L.L.; Zhang, B.; Yang, X.L.; Zhang, S.S.; Wang, Y.; Wang, G.D. Sensing platform of PdO-ZnO-In2 O3 nanofibers using MOF
templated catalysts for triethylamine detection. Sens. Actuators B 2021, 343, 130126. [CrossRef]
6. Li, S.H.; Xie, L.L.; He, M.; Hu, X.B.; Luo, G.F.; Chen, C.; Zhu, Z. Metal-Organic frameworks-derived bamboo-like CuO/In2 O3
Heterostructure for high-performance H2 S gas sensor with Low operating temperature. Sens. Actuators B 2020, 310, 127828.
[CrossRef]
7. Ueda, T.; Boehme, I.; Hyodo, T.; Shimizu, Y.; Weimar, U.; Barsan, N. Enhanced NO2 sensing properties of Au-loaded porous
In2 O3 gas sensors at low operating temperatures. Chemosensors 2020, 8, 72. [CrossRef]
8. Yang, X.L.; Zhang, S.F.; Yu, Q.; Zhao, L.P.; Sun, P.; Wang, T.S.; Liu, F.; Yan, X.; Gao, Y.; Liang, X.; et al. One step synthesis of
branched SnO2 /ZnO heterostructures and their enhanced gas-sensing properties. Sens. Actuators B 2019, 281, 415–423. [CrossRef]
9. Hiyoto, K.A.M.; Fisher, E.R. Utilizing plasma modified SnO2 paper gas sensors to better understand gas-surface interactions at
low temperatures. J. Vac. Sci. Technol. A 2020, 38, 043202. [CrossRef]
10. Zheng, L.; Bi, W.J.; Jin, Z.; Liu, S.T. Synthesis of hierarchical shell-core SnO2 microspheres and their gas sensing properties. Chin.
Chem. Lett. 2020, 31, 2083–2086. [CrossRef]
11. Ueda, T.; Maeda, T.; Huang, Z.D.; Higuchi, K.; Izawa, K.; Kamada, K.; Hyodo, T.; Shimizu, Y. Enhancement of methylmercaptan
sensing response of WO3 semiconductor gas sensors by gas reactivity and gas diffusivity. Sens. Actuators B 2018, 273, 826–833.
[CrossRef]
12. Buyukkose, S. Highly selective and sensitive WO3 nanoflakes based ammonia sensor. Mater. Sci. Semicond. Process. 2020,
110, 104969. [CrossRef]
13. Wang, C.Y.; Li, Y.H.; Qiu, P.P.; Duan, L.L.; Bi, W.; Chen, Y.; Guo, D.; Liu, Y.; Luo, W.; Deng, Y. Controllable synthesis of highly
crystallized mesoporous TiO2 /WO3 heterojunctions for acetone gas sensing. Chin. Chem. Lett. 2020, 31, 1119–1123. [CrossRef]
14. Mokoena, T.P.; Tshabalala, Z.P.; Hillie, K.T.; Swart, H.C.; Motaung, D.E. The blue luminescence of p-type NiO nanostructured
material induced by defects: H2 S gas sensing characteristics at a relatively low operating temperature. Appl. Surf. Sci. 2020,
525, 146002. [CrossRef]
15. Simion, C.E.; Ghica, C.; Mihalcea, C.G.; Ghica, D.; Mercioniu, I.; Somacescu, S.; Florea, Q.G.; Stanoiu, A. Insights about CO
gas-sensing mechanism with NiO-based gas sensors the influence of humidity. Chemosensors 2021, 9, 244. [CrossRef]
16. Alev, O.; Sarica, N.; Ozdemir, O.; Arslan, L.C.; Buyukkose, S.; Oztur, Z.Z. Cu-doped ZnO nanorods based QCM sensor for
hazardous gases. J. Alloys Compd. 2020, 826, 154177. [CrossRef]
17. Umar, A.; Algadi, H.; Kumar, R.; Akhtar, M.S.; Ibrahim, A.A.; Albargi, H.; Alhamami, M.A.M.; Alsuwian, T.; Zeng, W. Ultrathin
leaf-shaped CuO nanosheets based sensor device for enhanced hydrogen sulfide gas sensing application. Chemosensors 2021,
9, 221. [CrossRef]
18. Reynolds, D.C.; Look, D.C.; Jogai, B.; Hoelscher, J.E.; Sherriff, R.E.; Harris, M.T.; Callahan, M.J. Time-resolved photoluminescence
lifetime measurements of the Γ5 and Γ6 free excitons in ZnO. J. Appl. Phys. 2000, 88, 2152–2153. [CrossRef]
Chemosensors 2022, 10, 327 14 of 15

19. Norton, D.P.; Heo, Y.W.; Ivill, M.P.; Ip, K.; Pearton, S.J.; Chisholm, M.F. ZnO: Growth, doping & processing. Mater. Today 2004,
7, 34–40.
20. Modaberi, M.R.; Rooydell, R.; Brahma, S.; Akande, A.A.; Mwakikunga, B.W.; Liu, C.P. Enhanced response and selectivity of H2 S
sensing through controlled Ni doping into ZnO nanorods by using single metal organic precursors. Sens. Acuators B 2018, 273,
1278–1290. [CrossRef]
21. Liang, Y.C.; Liao, W.K.; Deng, X.S. Synthesis and substantially enhanced gas sensing sensitivity of homogeneously nanoscale Pd-
and Au-particle decorated ZnO nanostructures. J. Alloys Compd. 2014, 599, 87–92. [CrossRef]
22. Gardon, M.; Guilemany, J.M. A review on fabrication, sensing mechanisms and performance of metal oxide gas sensors. J. Mater.
Sci. Mater. Electron. 2013, 24, 1410–1421. [CrossRef]
23. Hu, J.; Gao, F.Q.; Sang, S.B.; Li, P.W.; Deng, X.; Zhang, W.D.; Chen, Y.; Lian, K. Optimization of Pd content in ZnO microstructures
for high-performance gas detection. J. Mater. Sci. 2015, 50, 1935–1942. [CrossRef]
24. Wang, W.C.; Tian, Y.T.; Wang, X.C.; He, H.; Xu, Y.R.; He, C.; Li, H. Ethanol sensing properties of porous ZnO spheres via
hydrothermal route. J. Mater. Sci. 2013, 48, 3232–3238. [CrossRef]
25. Khoang, N.D.; Hong, H.S.; Trung, D.D.; van Duy, N.; Hoa, N.D.; Thinh, D.D.; Van Hieu, N. On-chip growth of wafer-scale
planar-type ZnO nanorod sensors for effective detection of CO gas. Sens. Actuators B 2013, 181, 529–536. [CrossRef]
26. Zhang, W.H.; Zhang, W.D.; Zhou, J.F. Solvent thermal synthesis and gas-sensing properties of Fe-doped ZnO. J. Mater. Sci. 2010,
45, 209–215. [CrossRef]
27. Luo, J.; Ma, S.Y.; Li, F.M.; Li, X.B.; Li, W.Q.; Cheng, L.; Mao, Y.Z.; Gz, D.J. The mesoscopic structure of flower-like ZnO nanorods
for acetone detection. Mater. Lett. 2014, 121, 137–140. [CrossRef]
28. Xiong, H.M. ZnO nanoparticles applied to bioimaging and drug delivery. Adv. Mater. 2013, 25, 5329–5335. [CrossRef]
29. Huang, J.R.; Xu, X.J.; Gu, C.P.; Yang, M.; Yang, M.; Liu, J.H. Large-scale synthesis of hydrated tungsten oxide 3D architectures by
a simple chemical solution route and their gas-sensing properties. J. Mater. Chem. 2011, 21, 13283–13289. [CrossRef]
30. Li, H.Y.; Wang, X. Three-dimensional architectures constructed using two-dimensional nanosheets. Sci. China Chem. 2015, 58,
1792–1799. [CrossRef]
31. Sohila, S.; Rajendran, R.; Yaakob, Z.; Teridi, M.A.M.; Sopian, K. Photoelectrochemical water splitting performance of flower like
ZnO nanostructures synthesized by a novel chemical method. J. Mater. Sci. Mater. Electron. 2016, 27, 2846–2851. [CrossRef]
32. Akamatsu, T.; Itoh, T.; Tsuruta, A.; Masuda, Y. CH3 SH and H2 S Sensing Properties of V2 O5 /WO3 /TiO2 Gas Sensor. Chemosensors
2021, 9, 113. [CrossRef]
33. Li, Y.M.; Hu, H.J.; Zhang, W.F.; Tian, Z.Q.; Jiang, X.Q.; Wang, Y.H.; Zhang, S.; Zhang, Q.; Jian, J.; Zou, J. Theoretical Study on the
Electrochemical Catalytic Activity of Au-Doped Pt Electrode for Nitrogen Monoxide. Chemosensors 2022, 10, 178. [CrossRef]
34. Guo, W.W.; Jian, L.J.; Wang, X.M.; Zeng, W. Hydrothermal synthesis of Ni-doped hydrangea-like Bi2 WO6 and the enhanced gas
sensing property to n-butanol. Sens. Actuators B 2022, 357, 131396. [CrossRef]
35. Sun, Q.H.; Wu, Z.F.; Cao, B.B.; Chen, X.; Zhang, C.C.; Shaymurat, T.; Duan, H.; Zhang, J.; Zhang, M. Gas sensing performance of
biomass carbon materials promoted by nitrogen doping and p-n junction. Appl. Surf. Sci. 2022, 592, 153254. [CrossRef]
36. Park, H.; Kim, D.H.; Ma, B.S.; Shin, E.; Kim, Y.; Kim, T.S.; Kim, F.S.; Kim, I.D.; Kim, B.J. High-Performance, Flexible NO2
Chemiresistors Achieved by Design of Imine-Incorporated n-Type Conjugated Polymers. Adv. Sci. 2022, 9, 2200270. [CrossRef]
37. Fan, Y.R.; Xu, Y.Y.; Wang, Y.X.; Sun, Y.Q. Fabrication and characterization of Co-doped ZnO nanodiscs for selective TEA sensor
applications with high response, high selectivity and ppb-level detection limit. J. Alloys Compd. 2021, 876, 160170. [CrossRef]
38. Lu, S.H.; Hu, X.F.; Zheng, H.; Qiu, J.W.; Tian, R.B.; Quan, W.J.; Min, X.; Ji, P.; Hu, Y.; Cheng, S.; et al. Highly selective, ppb-level
xylene gas detection by Sn2+ -doped NiO flower-like microspheres prepared by a one-step hydrothermal. Method Sens. 2019,
19, 2958. [CrossRef]
39. Lee, C.S.; Li, H.Y.; Kim, B.Y.; Jo, Y.M.; Byun, H.G.; Hwang, I.S.; Abdel-Hady, F.; Wazzan, A.A.; Lee, J.-H. Discriminative detection
of indoor volatile organic compounds using a sensor array based on pure and Fe-doped In2 O3 nanofibers. Sens. Actuators B 2019,
285, 193–200. [CrossRef]
40. Chacko, L.; Massera, E.; Aneesh, P.M. Enhancement in the selectivity and sensitivity of Ni and Pd functionalized MoS2 toxic gas
sensors. J. Electrochem. Soc. 2020, 167, 106506. [CrossRef]
41. Guo, W.W.; Zhao, B.Y.; Zhou, Q.L.; He, Y.Z.; Wang, Z.C.; Radacsi, N. Fe-doped ZnO/reduced graphene oxide nanocomposite
with synergic enhanced gas sensing performance for the effective detection of formaldehyde. ACS Omega 2019, 4, 10252–10262.
[CrossRef]
42. Lu, Y.Y.; Zhan, W.W.; He, Y.; Wang, Y.T.; Kong, X.J.; Kuang, Q.; Xie, Z.; Zheng, X. MOF-templated synthesis of porous Co3 O4
concave nanocubes with high specific surface area and their gas sensing properties. ACS Appl. Mater. Interfaces 2014, 6, 4186–4195.
[CrossRef]
43. Li, D.P.; Zhang, Y.; Xu, J.C.; Jin, H.X.; Jin, D.F.; Hong, B.; Peng, X.; Wang, P.; Ge, H.; Wang, X. Nanocasting synthesis and
gas-sensing behavior of hematite nanowires. Phys. E Low-Dimens. Syst. Nanostruct. (Amst. Neth.) 2016, 84, 395–400. [CrossRef]
44. Chen, H.D.; Jin, K.L.; Wang, P.F.; Xu, J.C.; Han, Y.B.; Jin, H.X.; Jin, D.F.; Peng, X.L.; Hong, B.; Li, J.; et al. Highly enhanced
gas-sensing properties of indium-doped mesoporous hematite nanowires. J. Phys. Chem. Solids 2018, 120, 271–278. [CrossRef]
45. Su, C.; Zhang, L.; Han, Y.T.; Ren, C.; Li, B.L.; Wang, T.; Zeng, M.; Su, Y.; Hu, N.; Zhou, Z.; et al. Glucose-assisted synthesis
of hierarchical NiO-ZnO heterostructure with enhanced glycol gas sensing performance. Sens. Actuators B 2021, 329, 129167.
[CrossRef]
Chemosensors 2022, 10, 327 15 of 15

46. Sun, K.; Zhan, G.H.; Chen, H.D.; Lin, S.W. Low-operating-temperature NO2 sensor based on a CeO2 /ZnO heterojunction. Sensors
2021, 21, 8269. [CrossRef] [PubMed]
47. Zhao, Y.M.; Wang, S.; Zhai, X.; Shao, L.; Bai, X.J.; Liu, Y.L.; Wang, T.; Li, Y.; Zhang, L.; Fan, F.; et al. Construction of Zn/Ni
bimetallic organic framework derived ZnO/NiO heterostructure with superior N-propanol sensing performance. ACS Appl.
Mater. Interfaces 2021, 13, 9206–9215. [CrossRef] [PubMed]
48. Duan, Z.H.; Zhao, Q.N.; Wang, S.; Huang, Q.; Yuan, Z.; Zhang, Y.J.; Jiang, Y.; Tai, H. Halloysite nanotubes: Natural, environmental-
friendly and low-cost nanomaterials for high-performance humidity sensor. Sens. Actuators B 2020, 317, 128204. [CrossRef]
49. Chen, X.X.; Shen, Y.B.; Zhou, P.F.; Zhong, X.X.; Li, G.D.; Han, C.; Wei, D.; Li, S. Bimetallic Au/Pd nanoparticles decorated ZnO
nanowires for NO2 detection. Sens. Actuators B 2019, 289, 160–168. [CrossRef]
50. Zhou, C.G.; Meng, F.Q.; Chen, K.; Yang, X.L.; Wang, T.S.; Sun, P.; Liu, F.; Yan, X.; Shimanoe, K.; Lu, G. High sensitivity and low
detection limit of acetone sensor based on NiO/Zn2 SnO4 p-n heterojunction octahedrons. Sens. Actuators B 2021, 339, 129912.
[CrossRef]
51. Wang, C.; Cui, X.B.; Liu, J.Y.; Zhou, X.; Cheng, X.Y.; Sun, P.; Hu, X.; Li, X.; Zheng, J.; Lu, G. Design of superior ethanol gas sensor
based on Al-doped NiO nanorod-flowers. ACS Sens. 2016, 1, 131–136. [CrossRef]
52. Wang, S.; Huang, D.; Xu, S.S.; Jiang, W.K.; Wang, T.; Hu, J.; Hu, N.; Su, Y.; Zhang, Y.; Yang, Z. Two-dimensional NiO nanosheets
with enhanced room temperature NO2 sensing performance via Al doping. Phys. Chem. Chem. Phys. 2017, 19, 19043–19049.
[CrossRef]
53. Wang, L.L.; Lou, Z.; Zhang, R.; Zhou, T.T.; Deng, J.N.; Zhang, T. Hybrid Co3O4 /SnO2 core-shell nanospheres as real-time
rapid-response sensors for ammonia gas. ACS Appl. Mater. Interfaces 2016, 8, 6539–6545. [CrossRef]
54. Korotcenkov, G.; Cho, B.K. Engineering approaches for the improvement of conductometric gas sensor parameters Part 1.
Improvement of sensor sensitivity and selectivity (short survey). Sens. Actuators B 2013, 188, 709–728. [CrossRef]
55. Li, F.; Gao, X.; Wang, R.; Zhang, T. Design of WO3 -SnO2 core-shell nanofibers and their enhanced gas sensing performance based
on different work function. Appl. Surf. Sci. 2018, 442, 30–37. [CrossRef]
56. Kim, J.H.; Kim, H.W.; Kim, S.S. Self-heating effects on the toluene sensing of Pt-functionalized SnO2 -ZnO core-shell nanowires.
Sens. Actuators B 2017, 251, 781–794. [CrossRef]
57. Chen, H.; Ao, S.R.; Li, G.D.; Gao, Q.; Zou, X.X.; Wei, C.D. Enhanced sensing performance to toluene and xylene by constructing
NiGa2 O4 -NiO heterostructures. Sens. Actuators B 2019, 282, 331–338. [CrossRef]
58. Nie, L.F.; Fan, G.J.; Wang, A.Q.; Zhang, L.; Guan, J.; Han, N.; Chen, Y. Finely dispersed and highly toluene sensitive NiO/NiGa2 O4
heterostructures prepared from layered double hydroxides precursors. Sens. Actuators B 2021, 345, 130412. [CrossRef]
59. Hermawan, A.; Asakura, Y.; Inada, M.; Yin, S. A facile method for preparation of uniformly decorated-spherical SnO2 by CuO
nanoparticles for highly responsive toluene detection at high temperature. J. Mater. Sci. Technol. 2020, 51, 119–129. [CrossRef]
60. Jamnik, D.J.S.S.I. Investigation of the PTCR effect in ZnO–NiO two-phase ceramics. Solid State Ion. 1997, 99, 125–135.
61. Kim, H.J.; Lee, J.H. Highly sensitive and selective gas sensors using p-type oxide semiconductors: Overview. Sens. Actuators B
2014, 192, 607–627. [CrossRef]
62. Jun, J.H.; Yun, J.; Cho, K.; Hwang, I.S.; Lee, J.H.; Kim, S. Necked ZnO nanoparticle-based NO2 sensors with high and fast response.
Sens. Actuators B 2009, 140, 412–417. [CrossRef]
63. Zhou, J.Y.; Bai, J.L.; Zhao, H.; Yang, Z.Y.; Gu, X.Y.; Huang, B.Y.; Zhao, C.H.; Cairang, L.; Sun, G.Z.; Zhang, Z.X.; et al. Gas sensing
enhancing mechanism via doping-induced oxygen vacancies for gas sensors based on indium tin oxide nanotubes. Sens. Actuators
B 2018, 265, 273–284. [CrossRef]
64. Patil, V.L.; Vanalakar, S.A.; Patil, P.S.; Kim, J.H. Fabrication of nanostructured ZnO thin films based NO2 gas sensor via SILAR
technique. Sens. Actuators B 2017, 239, 1185–1193. [CrossRef]
65. Zhang, R.; Gao, S.; Zhou, T.T.; Tu, J.C.; Zhang, T. Facile preparation of hierarchical structure based on p-type Co3 O4 as toluene
detecting sensor. Appl. Surf. Sci. 2020, 503, 144167. [CrossRef]
66. Zhang, R.; Xu, Z.W.; Zhou, T.T.; Fei, T.; Wang, R.; Zhang, T. Improvement of gas sensing performance for tin dioxide sensor
through construction of nanostructures. JCIS 2019, 557, 673–682. [CrossRef]
67. Kim, J.H.; Mirzaei, A.; Kim, H.W.; Kim, S.S. Variation of shell thickness in ZnO-SnO2 core-shell nanowires for optimizing sensing
behaviors to CO, C6H6, and C7H8 gases. Sens. Actuators B 2020, 302, 127150. [CrossRef]
68. Bai, J.L.; Zhao, C.H.; Gong, H.M.; Wang, Q.; Huang, B.Y.; Sun, G.Z.; Wang, Y.; Zhou, J.; Xie, E.; Wang, F. Debye-length controlled
gas sensing performances in NiO@ZnO p-n junctional core-shell nanotubes. J. Phys. D Appl. Phys. 2019, 52, 285103. [CrossRef]
69. Zhou, T.T.; Zhang, T.; Zeng, Y.; Zhang, R.; Lou, Z.; Deng, J.N.; Wang, L. Structure-driven efficient NiFe2 O4 materials for ultra-fast
response electronic sensing platform. Sens. Actuators B 2018, 255, 1436–1444. [CrossRef]
70. Kim, H.R.; Choi, K.I.; Kim, K.M.; Kim, I.D.; Cao, G.Z.; Lee, J.H. Ultra-fast responding and recovering C2 H5 OH sensors using
SnO2 hollow spheres prepared and activated by Ni templates. Chem. Commun. 2010, 46, 5061–5063. [CrossRef]
71. Yang, M.; Lu, J.Y.; Wang, X.; Zhang, H.; Chen, F.; Sun, J.B.; Yang, J.; Sun, Y.; Lu, G. Acetone sensors with high stability to humidity
changes based on Ru-doped NiO flower-like microspheres. Sens. Actuators B 2020, 313, 127965. [CrossRef]
72. Luo, Y.B.; An, B.X.; Bai, J.L.; Wang, Y.R.; Cheng, X.; Wang, Q.; Li, J.; Yang, Y.; Wu, Z.; Xie, E. Ultrahigh-response hydrogen sensor
based on PdO/NiO co-doped In2 O3 nanotubes. J. Colloid Interface Sci. 2021, 599, 533–542. [CrossRef]

You might also like